You are on page 1of 13

Corrosion Science 53 (2011) 263–275

Contents lists available at ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Inhibitive properties, thermodynamic and quantum chemical studies of alloxazine


on mild steel corrosion in H2SO4
N.O. Obi-Egbedi a, I.B. Obot b,⇑
a
Department of Chemistry, University of Ibadan, Ibadan, Nigeria
b
Department of Chemistry, Faculty of Science, University of Uyo, P.M.B. 1017, Uyo, Akwa Ibom State, Nigeria

a r t i c l e i n f o a b s t r a c t

Article history: Alloxazine (ALLOX) was tested as corrosion inhibitor for mild steel in 0.5 M H2SO4 solution using non-
Received 30 April 2010 electrochemical technique (gravimetric and UV–Visible spectrophotometric measurements) at 303–
Accepted 9 September 2010 333 K. ALLOX acts as inhibitor for mild steel in acidic medium. Inhibition efficiency increases with
Available online 17 September 2010
increase in concentration of ALLOX but decrease with rise in temperature. The adsorption of ALLOX
was found to follow Temkin adsorption isotherm model. Both the activation and thermodynamic param-
Keywords: eters governing the adsorption process were calculated and discussed. The adsorption follows a first-
A. Alloxazine
order kinetics. DFT study gave further insight into the mechanism of inhibition action of ALLOX.
B. Mild steel
C. Sulphuric acid
Ó 2010 Elsevier Ltd. All rights reserved.
D. Density functional theory (DFT)

1. Introduction Lumichrome (7,8-dimethylalloxazine), for example was found to


inhibit flavin reductase in living Escherichia coli cells [17]. Another
Corrosion inhibitors have been widely used in stimulation oper- interesting application is in an optical transistor device with a thin
ations in petroleum wells [1,2]. In these operations acid solutions film of Lumichrome on conductive SnO2 glass [18]. A further point
at temperatures up to 333 K are often employed to remove iron of interest is the possibility of using alloxazines to sensitize the
oxides and carbonated minerals [3]. In such aggressive medium, photooxidation of substituted phenols in water [19]. However,
the use of corrosion inhibitors is one of the most common, effective there is no literature to date about the corrosion inhibitive effect
and economic methods to protect metals in acid media [4]. of alloxazine on mild steel in H2SO4 solution.
The majority of the well-known inhibitors are organic com- The aim of this paper therefore is to explore the use of alloxazine
pounds containing heteroatoms such as oxygen, nitrogen or sul- as an acid corrosion inhibitor for mild steel surface in sulphuric acid
phur, and multiple bonds, which allow an adsorption on the solution using gravimetric method. The effect of temperature on
metal surface [5–7]. It has been observed that the adsorption of corrosion and inhibition processes are thoroughly assessed and dis-
these inhibitors depends on the physico-chemical properties of cussed. Kinetic and thermodynamic parameters were calculated and
the functional groups and the electron density at the donor atom. discussed in detail. UV–Visible spectroscopy together with quantum
The adsorption occurs due to the interaction of the lone pair and/ chemical study using density functional theory were further em-
or p-orbitals of inhibitor with d-orbitals of the metal surface ployed to provide additional insight into the mechanism of inhibi-
atoms, which evokes a greater adsorption of the inhibitor mole- tory action.
cules onto the surface, leading to the formation of a corrosion pro-
tection film [8–10]. Furthermore, adsorption is also influenced by 2. Experimental method
the structure and the charge of metal surface, and the type of test-
ing electrolyte [11]. The choice of effective inhibitors is based on 2.1. Material
their mechanism of action and electron-donating ability. The sig-
nificant criteria involved in this selection are molecular structure, Test were performed on a freshly prepared sheet of mild steel of
electron density on the donor atoms, solubility and dispersibility the following composition (wt.%): 0.13% C, 0.18% Si, 0.39% Mn,
[12–15]. 0.40% P, 0.04% S, 0.025% Cu, and bal. Fe. Specimens used in the
Recently, interest in alloxazines has intensified because of their weight loss experiment were mechanically cut into 5.0 cm 
important role in a wide range of biological systems [16]. 4.0 cm  0.8 cm dimensions, then abraded with SiC abrasive
papers 320, 400 and 600 grit, respectively, washed in absolute eth-
⇑ Corresponding author. Tel.: +234 8067476065. anol and acetone, dried in room temperature and stored in a mois-
E-mail address: proffoime@yahoo.com (I.B. Obot). ture free desiccator before their use in corrosion studies [5,6].

0010-938X/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.corsci.2010.09.020
264 N.O. Obi-Egbedi, I.B. Obot / Corrosion Science 53 (2011) 263–275

ALLOX. It is evident that ALLOX is a heterocyclic compound con-


taining nitrogen and oxygen atoms, which could easily be proton-
ated in acidic solution, and several p-electrons exist in this
molecule.

2.3. Solutions

The aggressive solutions, 0.5 M H2SO4 were prepared by dilu-


Fig. 1. Molecular structure of alloxazine (ALLOX). tion of analytical grade 98% H2SO4 with distilled water. The con-
centration range of ALLOX was 2–10 lM.
2.2. Inhibitor
2.4. Gravimetric measurements
Alloxazine (ALLOX) was purchased from SIGMA–ALDRICH and
used as inhibitor. Stock solution was made in 10:1 water:methanol The gravimetric method (weight loss) is probably the most
mixture to ensure solubility [20]. This stock solution was used for widely used method of inhibition assessment [21–24]. The sim-
all experimental purposes. Fig. 1 shows the molecular structure of plicity and reliability of the measurement offered by the weight

Fig. 2. Variation of weight loss against time for mild steel corrosion in 0.5 M H2SO4 in the presence of different concentrations of ALLOX at (a) 303 (b) 313 (c) 323 and (d)
333 K.
N.O. Obi-Egbedi, I.B. Obot / Corrosion Science 53 (2011) 263–275 265

Fig. 2 (continued)

loss method is such that the technique forms the baseline method DW
q¼ ð1Þ
of measurement in many corrosion monitoring programmes [25]. St
Weight loss measurements were conducted under total immer-
where W is the average weight loss of three mild steel sheets, S the
sion using 250 mL capacity beakers containing 200 mL test solu-
total area of one mild steel specimen, and t is the immersion time
tion at 303–333 K maintained in a thermostated water bath.
(10 h). With the calculated corrosion rate, the inhibition efficiency
The mild steel coupons were weighed and suspended in the bea-
(%I) was calculated as follows [28,29]:
ker with the help of rod and hook. The coupons were retrieved at
 
2 h interval progressively for 10 h, washed thoroughly in 20% q1  q2
%I ¼  100 ð2Þ
NaOH solution containing 200 g/l of zinc dust [10] with bristle q1
brush, rinsed severally in deionized water, cleaned, dried in ace-
tone, and re-weighed. The weight loss, in grammes, was taken where q1 and q2 are the corrosion rates of the mild steel coupons in
as the difference in the weight of the mild steel coupons before the absence and presence of inhibitor, respectively.
and after immersion in different test solutions. Then the tests
were repeated at different temperatures. In order to get good 2.5. Spectrophotometric measurements
reproducibility, experiments were carried out in triplicate. In this
present study, the standard deviation values among parallel trip- UV–Visible absorption spectrophotometric method was carried
licate experiments were found to be smaller than 5%, indicating out on the prepared mild steel samples after immersion in 0.5 M
good reproducibility. H2SO4 with and without addition of 10 lM of alloxazine at 303 K
The corrosion rate (q) in mg cm2 h1 was calculated from the for 3 days. All the spectra measurements were carried out using
following equation [26,27]: a Perkin-Elmer UV–Visible Lambda 2 spectrophotometer.
266 N.O. Obi-Egbedi, I.B. Obot / Corrosion Science 53 (2011) 263–275

θ
Fig. 5. Variation of the degree of surface coverage (h) with the logarithm of bulk
concentration of ALLOX.

Fig. 3. The relationship between corrosion rate and temperature in the absence and
presence of different concentrations of ALLOX. (d,p) level of theory using Spartan’06 V112 program package [31]
in order to determine whether they correspond to a maximum or
a minimum in the potential energy curve. The theoretical parame-
2.6. Computational details ters were calculated for molecules in gas phase, aqueous phase and
in protonated form for comparison. It is well known that the phe-
B3LYP, a version of the DFT method that uses Becke’s three nomenon of electrochemical corrosion occurs in liquid phase. As a
parameter functional (B3) and includes a mixture of HF with DFT result, it was necessary to include the effect of a solvent in the
exchange terms associated with the gradient corrected correlation computational calculations. In the Spartan’06 V112 program, SCRF
functional of Lee, Yang and Parr (LYP) [30], was used in this paper methods (Self-consistent reaction field) were used to perform cal-
to carry out quantum calculations. Then, full geometry optimiza- culations in solution. These methods model the solvent as a contin-
tion together with the vibrational analysis of the optimized struc- uum of uniform dielectric constant and the solute is placed in the
tures of the inhibitor was carried out at the highest (B3LYP/6-31G cavity within it.

Table 1
Calculated values of corrosion rate and inhibition efficiency for mild steel corrosion for mild steel in 0.5 M H2SO4 in the absence and presence of ALLOX at 303–333 K.

System/concentration Corrosion rate (mg cm2 h1) Inhibition efficiency (%I)


303 K 313 K 323 K 333 K 303 K 313 K 323 K 333 K
Blank 0.99 1.84 2.84 4.40 – – – –
2 lM 0.89 1.68 2.67 4.14 10 09 07 06
4 lM 0.50 1.08 1.07 3.56 49 41 35 20
6 lM 0.24 0.55 1.19 3.42 76 70 59 22
8 lM 0.11 0.33 1.01 2.62 89 82 65 40
10 lM 0.03 0.23 0.70 2.22 96 88 76 50

Fig. 4. Variation of inhibition efficiency of ALLOX with temperature.


N.O. Obi-Egbedi, I.B. Obot / Corrosion Science 53 (2011) 263–275 267

3. Results and discussion 3.2. Effect of temperature

3.1. Effect of alloxazine on weight loss, corrosion rate and inhibition Temperature has a great effect on the rate of metal electro-
efficiency chemical corrosion. In case of corrosion in a neutral solution (oxy-
gen depolarization) the increase in temperature has a favourable
Corrosion inhibition performance of organic compounds as cor- effect on the overpotential of oxygen depolarization and the rate
rosion inhibitors can be evaluated using electrochemical and of oxygen diffusion but it leads to a decrease of oxygen solubility.
chemical techniques. For the chemical methods, a weight loss mea- In case of corrosion in acidic medium (hydrogen depolarization),
surement is ideally suited for long term immersion test. Corrobora- the corrosion rate increases exponentially with temperature
tive results between weight loss and other techniques have been increase because the hydrogen evolution overpotential decreases
reported [32–35]. [39]. The relationship between the corrosion rate (q) of mild steel
The anodic dissolution of iron in acidic media and the corre- in acidic media and temperature (T) is often expressed by the
sponding cathodic reaction has been reported to proceed as follows Arrhenius equation [40,41]:
[27]:
Ea
log q ¼ log A  ð5Þ
2þ 2:303RT
Fe ! Fe þ 2e ð3Þ
2Hþ þ 2e ! 2Hads ! H2 ð4Þ where q is the corrosion rate Ea is the apparent activation energy, R
is the molar gas constant (8.314 J K1 mol1), T is the absolute tem-
As a result of these reactions, including the high solubility of the perature, and A is the frequency factor. The plot of log q against 1/T
corrosion products, the metal loses weight in the solution. Fig. 2a–d for mild steel corrosion in 0.5 M H2SO4 in the absence and presence
shows the plots of weight loss versus time for mild steel in 0.5 M of different concentrations of ALLOX is presented in Fig. 6. All
H2SO4 without and with different concentrations of ALLOX at 303– parameters were given in Table 2.
333 K, respectively. From the plots, it is evident that the weight loss The activation energy increased in the presence of ALLOX,
of mild steel in the different test solutions increases with time. The which indicated physical (electrostatic) adsorption [42]. Further-
non-uniformity and non-linearity of the curves of the weight loss more, the activation energy rose with increasing inhibitor concen-
plot may be attributed to the presence of mill scale on the mild steel tration, suggesting strong adsorption of inhibitor molecules at the
surface. It may also suggest that the mild steel corrosion by H2SO4 is metal surface [43]. The increase in activation energy was due to the
a heterogeneous process involving several steps. Similar observa- corrosion reaction mechanism in which charge transfer was
tion has been reported recently [27]. A further inspection of the plots blocked by the adsorption of ALLOX molecules on the mild steel
reveal that the weight loss of mild steel was reduced in the presence surface [44,45]. It also revealed that the whole process was con-
of ALLOX compared to the free acid solution; an indication of inhib- trolled by the surface reaction since the energy of the activation
iting effect of acid corrosion of mild steel. corrosion process in both the absence and presence of ALLOX
A comparison of corrosion rate of mild steel with temperature was greater than 20 kJ mol1 [46].
(303–333 K) in the absence and presence of different concentra- Experimental corrosion rate values obtained from weight loss
tions of ALLOX are dipicted in Fig. 3 and the calculated values pre- measurements for mild in 0.5 M H2SO4 in the absence and pres-
sented in Table 1. From Table 1 and Fig. 3, it is clear that corrosion ence of ALLOX was used to further gain insight on the change of en-
rate in the presence of ALLOX increases with rise in temperature. thalpy (DH ) and entropy (DS ) of activation for the formation of
This may be due to the fact that corrosion of mild steel in acidic the activation complex in the transition state using transition
environments is generally accompanied with hydrogen gas evolu- equation [47]:
tion and increase in temperature usually accelerates the evolution      
of hydrogen which results in higher dissolution rate of the mild
RT DS DH
q¼ exp exp ð6Þ
steel [20]. Fig. 3 also reveals that the corrosion rate was reduced Nh R RT
in the presence of different concentrations of ALLOX with the low- where q is the corrosion rate, h is the Plank’s constant (6.626176 
est value obtained at the highest concentration (10 lM) of ALLOX 1034 Js), N is the Avogadro’s number (6.02252  1023 mol1), R is
used at all the temperatures studied.
The influence of temperature on percentage inhibition effi-
ciency was studied by conducting weight loss measurements at
303–333 K containing different concentrations of ALLOX (Table
1). Fig. 4 shows the variation of percentage inhibition efficiency
with temperature. It is clear from the figure that percentage inhi-
bition efficiency increases with concentration but decreases with
temperature. The increase in percentage inhibition efficiency of
ALLOX with concentration may be due to the adsorption of allox-
azine onto the mild steel surface though non-bonding electron
pairs present on nitrogen and oxygen atoms as well as p-electrons
[36]. The decrease in inhibition efficiency with increase in temper-
ρ

ature may be probably due to increased rate of desorption of


ALLOX from the mild steel surface at higher temperature [37].
The variation of the degree of surface coverage (h) with the log-
arithm of bulk concentration is presented in Fig. 5 for the corrosion
medium- ALLOX systems at 303–333 K. The pattern of the graphs
in Fig. 5 describes an S-shaped curve [38]. These indicate the for-
mation of a protective barrier film (adsorption) of inhibitor mole-
cules on mild steel surface. This implies that ALLOX act as
inhibitor by establishment of a thin film which acts as a barrier Fig. 6. Arrhenius plot for mild steel corrosion in 0.5 M H2SO4 in the absence and
to the transport of the metal ions from the metal to the solution. presence of different concentrations of alloxazine (ALLOX).
268 N.O. Obi-Egbedi, I.B. Obot / Corrosion Science 53 (2011) 263–275

Table 2 charge sharing or charge transfer from the inhibitor molecules to


Activation parameters of the dissolution of mild steel in 0.5 M H2SO4 in the absence the metal surface to form a coordinate type bond. In fact electron
and presence of different concentrations of ALLOX.
transfer is typically for transition metals having vacant low-energy
Conc. A (g cm2 h1) Ea (kJ mol1) DH* (kJ mol1) DS* (J mol1 K1) electron orbital. Chemisorption is typified by much stronger
Blank 1.18  104
40.99 180.07 207.97 adsorption energy than physical adsorption. Such a bond is there-
2 lM 1.78  104 42.29 184.86 220.80 fore more stable at higher temperatures.
4 lM 8.60  105 53.54 223.25 331.85 Basic information on the adsorption of inhibitor on metal sur-
6 lM 6.59  108 72.35 286.06 516.24
8 lM 1.77  1011 88.34 340.24 675.35
faces can be provided by adsorption isotherm. Attempts were
10 lM 5.62  1015 117.16 437.32 966.00 made to fit experimental data to various isotherms including
Frumkin, Langmuir, Temkin, Freundlich, Bockris-Swinkels and Flo-
ry-Huggins isotherms. All these isotherms are of the general form
[50]:
the universal gas constant and T is the absolute temperature. Fig. 7
shows the plot of log q/T versus 1/T for mild steel corrosion in f ðh; xÞ expð2ahÞ ¼ K ads C ð7Þ
0.5 M H2SO4 in the absence and presence of different concentrations
where f ðh; xÞ is the configurational factor which depends on the
of ALLOX. Straight lines were obtained with slope of (DH /2.303R)
physical mode and the assumptions underlying the derivation of
and an intercept of [log (R/Nh) + (DS /2.303R)] from which the values
the isotherm, h the degree of surface coverage, C the inhibitor con-
of DH and DS , respectively, were computed and listed also in Table
centration, x the size factor ratio, a the molecular interaction
2. Inspection of these data reveal that the activation parameters (DH
parameter, and Kads the equilibrium constant of the adsorption pro-
and DS ) of dissolution reaction of mild steel in 0.5 M H2SO4 in the
cess. In this study, correlation coefficient (R2) was used to deter-
presence of ALLOX are higher than in the absence of inhibitor. The po-
mine the best fit isotherm which was obtained from Temkin
sitive sign of the enthalpy of activation reflect the endothermic nat-
adsorption isotherm. According to this isotherm, h is related to
ure of steel dissolution process meaning that dissolution of steel is
the inhibitor concentration by the following equation [51]:
difficult [48]. The entropy of activation was also positive in the ab-
sence and presence of ALLOX implying that the rate-determining step expð2ahÞ ¼ K ads C ð8Þ
for the activated complex is dissociation step rather than association.
where the molecular interaction parameter a can have both positive
In other words, the adsorption process is accompanied by an increase
and negative values. Positive values of a indicate attraction forces
in entropy, which is the driving force for the adsorption of inhibitor
between the adsorbed molecules while negative values indicate
onto the mild steel surface [49].
repulsive forces between the adsorbed molecules [51]. Upon rear-
rangement of Eq. (8), the following equation is obtained:
3.3. Adsorption isotherm and thermodynamic consideration
h ¼ ½1=ð2aÞ lnðK ads CÞ ð9Þ
Generally speaking, corrosion inhibitors are found to protect If the parameter f is defined as:
steel corrosion in acid solutions by adsorbing themselves on steel
surface. Moreover, the adsorption process depends on the mole- f ¼ 2a ð10Þ
cule’s chemical composition, the temperature and the electro- where f is the heterogeneous factor of the metal surface describing
chemical potentials at the metal/solution interface. Adsorption is the molecular interactions in the adsorption layer and the heteroge-
a separation process involving two phases between which certain neity of the metal surface. Eq. (10) clearly shows that the sign be-
components can be described by two main types of interaction tween f and a is reverse, that is, if a < 0, then f > 0; if a > 0, then
[47]: (1) physisorption which involves electrostatic forces between f < 0. Accordingly, if f > 0, mutual repulsion of molecules occurs
ionic charges at the metal/solution interface. The heat of adsorp- and if f < 0 attraction takes place. If Eq. (10) is substituted into Eq.
tion is low and therefore this type of adsorption is stable only at (9), then the Temkin isotherm equation [52] has the following form:
relatively low temperatures and; (2) chemisorption which involves
h ¼ ð1=f Þ lnðK ads CÞ ð11Þ
where h is the degree of surface coverage, and could be calculated
by the following relationship [27]:
%I
h¼ ð12Þ
100
Eq. (11) can be transformed into:
h ¼ ð1=f Þ ln K ads þ ð1=f Þ ln C ð13Þ
Eq. (13) is a different form of the Temkin isotherm. The plot of h ver-
sus lnC gives a straight line graph with a slope of (1/f) and an inter-
cept of [(1/f)lnKads]. It is clear that f can be calculated from the slope,
with the calculated f, the value of Kads can be obtained from the
intercept. All the parameters are listed in Table 3. Fig. 8, is the
ρ

Table 3
Some parameters from Temkin isotherm model for mild steel in 0.5 M H2SO4.

Temperature (K) f Kads (M1) (R2) DGoads (kJ mol1)


303 6.09 1.68 0.987 10.32
313 7.04 1.17 0.991 10.14
323 9.61 0.94 0.941 10.25
Fig. 7. Transition state plot for mild steel corrosion in 0.5 M H2SO4 in the absence 333 10.86 0.42 0.900 9.88
and presence of different concentrations of alloxazine (ALLOX).
N.O. Obi-Egbedi, I.B. Obot / Corrosion Science 53 (2011) 263–275 269

Fig. 9 shows the plot of lnKads versus 1/T which gives straight lines
with slope of (DHoads /R) and intercepts of (DSoads /R + ln 1/55.5).
Calculated value of DHoads and DSoads using the Van’t Hoff equation
are 13.60 kJ mol1 and 0.012 J mol1 K1, respectively.
The enthalpy and entropy for the adsorption of ALLOX on mild
steel were also deduced from the thermodynamic basic equation
[63]:

DGoads ¼ DHoads  T DSoads ð16Þ


θ

where DHoads and DSoads are the enthalpy and entropy changes of
adsorption process, respectively. A plot of DGoads versus T was linear
(Fig. 10) with the slope equal to DSoads and intercept of DHoads . The
enthalpy of adsorption DHoads , and the entropy of adsorption DSoads ,
obtained are 13.99 kJ mol1 and 0.018 J mol1 K1, respectively.
The negative sign of DHoads indicates that the adsorption of ALLOX
molecules is an exothermic process. In an exothermic process, phys-
isorption is distinguished from chemisorption by considering the
absolute value of DHoads . For physisorption process, the enthalpy of
adsorption is lower than 40 kJ mol1 while that for chemisorption
Fig. 8. The relationship between h and lnC at different temperatures. approaches 100 kJ mol1 [64]. In the present case, DHoads is lower
than 40 kJ mol1 clearly indicating that physical adsorption is in-
volved. Values of DHoads obtained by the two methods are in good
agreement.
relationship between h and lnC at different temperatures. These re-
The entropy of adsorption obtained from Eqs. (15) and (16) are
sults show that all the linear correlation coefficients (R2) are close to
negative because inhibitor molecule freely moving in the bulk solu-
1, which indicates that the adsorption of ALLOX onto steel surface
tion (inhibitor molecule were chaotic), were adsorbed in an orderly
obeys the Temkin adsorption isotherm. Furthermore, it can be de-
fashion onto the mild steel, resulting in a decrease in entropy [65].
duced that there is repulsion force in the adsorption layer due to
Moreover, from thermodynamic principles, since the adsorption
f > 0. Table 3 also indicates that the adsorption equilibrium constant
Kads values decrease with increasing temperature. Large values of
Kads mean better inhibition efficiency of the inhibitor, i.e., strong
electrical interaction between the double-layer existing at the
phase boundary and the adsorbing inhibitor molecules. Small val-
ues of Kads, however, reveal that such interactions between adsorb-
ing inhibitor molecules and the metal surface are weaker, indicating
that the inhibitor molecules are easily removable by the solvent
molecules from the metal surface [53]. These results confirm the
suggestion that ALLOX is physically adsorbed on the metal surface
and that the strength of the adsorption decreases with temperature.
Kads is related to the free energy of adsorption DGoads by the equa-
tion [54]:

DGoads
log K ads ¼  log C H2 O  ð14Þ
2:303RT
where C H2 O is the concentration of water expressed in mol/L (the
same as that of inhibitor concentration), R is the molar gas constant
(kJ mol1 K1) and T is the absolute temperature (K). Fig. 9. The relationship between lnKads and 1/T.
Calculated free energies DGoads values are given also in Table 3.
The negative values of DGoads indicate spontaneous adsorption of
ALLOX onto the mild steel surface [55] and strong interactions be-
tween inhibitor molecules and the metal surface [56]. Generally,
values of DGoads up to 20 kJ mol1 are consistent with physisorp-
tion, while those around 40 kJ mol1 or higher are associated
with chemisorption as a result of the sharing or transfer of elec-
trons from organic molecules to the metal surface to form a coor-
dinate bond [57–61]. The calculated values of DGoads are less than
20 kJ mol1, indicating that the adsorption mechanism of ALLOX
on mild steel in 0.5 M H2SO4 solution at the studied temperatures
Δ

is physisorption [43].
Thermodynamic parameters such as enthalpy of adsorption
DHoads and entropy of adsorption DSoads can be deduced from inte-
grated version of the Van’t Hoff equation expressed by [62]:

DHoads DSoads 1
ln K ads ¼  þ þ ln ð15Þ
RT R 55:5 Fig. 10. The relationship between DGoads and temperature.
270 N.O. Obi-Egbedi, I.B. Obot / Corrosion Science 53 (2011) 263–275

Fig. 11. The plot of lnWt against t for mild steel corrosion in 0.5 M H2SO4 with and without ALLOX.

was an exothermic process, it must be accompanied by a decrease 0:693


t1=2 ¼ ð23Þ
in entropy [66]. k
The values of rate constant and half-life periods obtained are sum-
3.4. Kinetics of mild steel corrosion in H2SO4 with and without
marized in Table 4. The value of the rate constant (k) was found to
alloxazine
be higher in the case of inhibited mild steel samples than uninhib-
ited samples whereas half-life (t1/2) was found to be lower in the
The kinetics of the mild steel corrosion in the absence and pres-
presence of inhibitor than in its absence. This implies that the cor-
ence of different concentrations of ALLOX in 0.5 M H2SO4 was stud-
rosion rate is higher in the absence of inhibitor than with inhibited
ied at 30 °C by fitting the corrosion data into different rate laws.
Correlation coefficients R2 were used to determine the best rate
law for the corrosion process. The rate laws considered were [67]:
Table 4
Zero-order : W t ¼ kt ð17Þ The values of first-order rate constant, half-life and time constant for the dissolution
of mild steel in 0.5 M H2SO4 in the absence and presence of different concentrations of
First-order : ln W t ¼ kt þ ln W o ð18Þ ALLOX.

Conc. k (h1) t1/2 (h) s (h)


Second-order : 1=W t ¼ kt þ 1=W o ð19Þ
Blank 0.216 3.20 4.63
where Wo is the initial weight of mild steel, Wt is the weight loss of 2 lM 0.307 2.25 3.25
mild steel at time t and k is the rate constant. 4 lM 0.292 2.37 3.42
6 lM 0.388 1.78 2.57
By far best result was obtained for first-order kinetics. The plot 8 lM 0.286 2.42 3.49
of lnWt against t which was linear (Fig. 11) with good correlation 10 lM 1.043 0.66 0.95
coefficients (R2 > 0.9) confirms first-order kinetics for the corrosion
of mild steel in 0.5 M H2SO4 in the absence and presence of ALLOX.
The corrosion rate of mild steel in sulphuric acid medium is un-
der anodic control [68] which is:
rds
Fe þ OH ¡ FeOHads þ Hþ þ e ð20Þ
þ
FeOHads ! FeOH þ e ð21Þ

FeOHþ þ Hþ ¡ Fe2þ þ H2 O ð22Þ

where ‘rds’ stands for rate-determining step. a


Fig. 11 reflects the reaction order with respect to mild steel. This
b
result suggests that the presence of ALLOX does not influence the
anodic reaction order. The linearity of the curves in the absence
and presence of ALLOX implies that its presence does not change
the kinetics of the corrosion reaction though the rate might be con-
siderably reduced. Similar report has been documented elsewhere
[69].
The half-life (t1/2) value was calculated [70] using the following Fig. 12. UV–Visible spectra of the solution containing 0.5 M H2SO4 (10 lM) ALLOX
relationship: (a) before and (b) after 3 days of mild steel immersion.
N.O. Obi-Egbedi, I.B. Obot / Corrosion Science 53 (2011) 263–275 271

mild steel samples. Similar results were reported on the kinetics of In order to confirm the possibility of the formation of ALLOX-Fe
some organic compounds [71]. complex, UV–Visible absorption spectra obtained from 0.5 M
Another indication of the rate of a first-order reaction is the H2SO4 solution containing 10 lM ALLOX before and after 3 days
time constant, s, [71]. The time required for the concentration of of mild steel immersions are shown in Fig. 12.
the reactant to fall to 1/e of its initial value. The calculated values The absorption spectrum of the solution containing 10 lM
of the time constants are also presented in Table 4. The results ALLOX before the steel immersion (Fig. 12) shows an adsorption
show that the time constants were lower in the presence of ALLOX band in the range of 330–400 nm. It has been reported that allox-
than in its absence. This confirms that the presence of the inhibitor azine spectra in the near-UV have maxima at about 340–400 nm
decreases the dissolution of mild steel. [18]. The absorption band at this longer wavelength is as a result
of the presence of aromatic system in ALLOX which is highly con-
3.5. UV–Visible spectroscopic investigation jugated. Furthermore, the presence of lone pair of electrons on the
alloxazine nitrogens atoms and/or carbonyl group give rise to the
The adsorption of monochromatic light according to Abboud spectra around 330 nm due to n-p* transition. After 3 days of steel
et al. [72], is a suitable method for identification of complex ions. immersion (Fig. 12), it is clearly seen that the band shifted to a
The adsorption of light is proportional to the concentration of the longer wavelength (390–470 nm) and there is a decrease in absor-
adsorbing species. For routine analysis, a simple conventional tech- bance. It is clear that there was no significant difference in the
nique based on UV–Visible absorption is the most sensitive direct shape of the spectra before and after the immersion of ALLOX
spectrophotometric detection. Change in position of the absor- showing a possibility of weak interaction between ALLOX and mild
bance maximum and change in the value of absorbance indicate steel (physisorption). Alloxazines are multifunctional with numer-
the formation of a complex between two species in solution [73]. ous electron donor sites and has been reported as a potential ligand

Fig. 13. Optimized structure of alloxazine (ALLOX) (ball and stick model).

Fig. 14. The highest occupied molecular orbital (HOMO) density of alloxazine (ALLOX) using DFT at the B3LYP/6-31G (d,p) basis set level.
272 N.O. Obi-Egbedi, I.B. Obot / Corrosion Science 53 (2011) 263–275

for metal centres, which in the absence of additional coordination and to describe the structural nature of the inhibitor on the corro-
sites bind predominantly through N(4) and O(2) to form five-mem- sion process [80–82]. Furthermore, DFT is considered a very useful
bered, redox active a-iminoketo chelate rings [74]. These experi- technique to probe the inhibitor/surface interaction as well as to
mental findings give strong evidence for the possibility of the analyze the experimental data. Thus in the present investigation,
formation of a complex between Fe2+ cation and ALLOX in H2SO4. quantum chemical calculation using DFT was employed to explain
the experimental results obtained in this study and to further give
3.6. Quantum chemical studies insight into the inhibition action of ALLOX on the mild steel
surface.
The density functional theory (DFT) is one of the most impor- For the purpose of determining the active sites of the inhibitor
tant theoretical models used in explaining the science of solids molecule, three influence factors: natural atomic charge, distribu-
and chemistry. A number of chemical concepts have been corre- tion of frontier orbital, and Fukui indices are considered. According
lated within the framework of DFT [75]. The most fundamental to classical chemical theory, all chemical interactions are by either
parameter in DFT is the electron density q(r) upon which all the electrostatic or orbital. Electrical charges in the molecule were
chemical quantities are expressed [76]. The structural parameters obviously the driving force of electrostatic interactions it has been
calculated through the q(r) concept, compare well with the param- proven that local electron densities or charges are important in
eters calculated by the W concept [77]. Since the theory is simpler many chemical reactions and physico-chemical properties of com-
than quantum mechanics, the interest has grown in understanding pound [83]. Figs. 13–16 show the optimized geometry, the HOMO
the structure, properties, reactivity, and dynamics of atoms, mole- density distribution, the LUMO density distribution and the Mul-
cules and clusters using DFT. In the field of reaction chemistry, DFT liken charge population analysis plots for ALLOX molecule ob-
exceeds the limit of wave mechanics [78], and it is emerging as a tained with DFT at B3LYP/6-31G (d,p) level of theory.
unique approach for the study of reaction mechanism [79]. The reactive ability of the inhibitor is considered to be closely
Recently, the density functional theory (DFT) has been used to related to their frontier molecular orbitals, the HOMO and LUMO.
analyze the characteristics of the inhibitor/surface mechanism Higher HOMO energy (EHOMO) of the molecule means a higher

Fig. 15. The lowest unoccupied molecular orbital (LUMO) density of alloxazine (ALLOX) using DFT at the B3LYP/6-31G (d,p) basis set level.

Fig. 16. Mulliken charges population analysis of alloxazine (ALLOX) using DFT at the B3LYP/6-31G (d,p) basis set level.
N.O. Obi-Egbedi, I.B. Obot / Corrosion Science 53 (2011) 263–275 273

electron-donating ability to appropriate acceptor molecules with Table 4 shows that ALLOX in the protonated form has the lowest
low-energy empty molecular orbital and thus explains the adsorp- energy gap and lowest hardness; this agrees with the experimental
tion on metallic surfaces by way of delocalized pairs of p-electrons. results that ALLOX could have better inhibitive performance on
ELUMO, the energy of the lowest unoccupied molecular orbital signi- mild steel surface in the protonated form i.e. through electrostatic
fies the electron receiving tendency of a molecule. Accordingly, the interaction between the cation form of ALLOX and the vacant
difference between ELUMO and EHOMO energy levels (DE = ELUMO  d-orbital of mild steel (physisorption). This also agrees well with
EHOMO) and the dipole moment (l) were also determined. The the value of DGoads and DHoads obtained experimentally. The dipole
global hardness g is approximated as DE/2, and can be defined un- moment is another important electronic parameter that results
der the principle of chemical hardness and softness (HSAB) [84]. from non-uniform distribution of charges on the various atoms in
These parameters also provide information about the reactive a molecule. It is mainly used to study the intermolecular interac-
behavior of molecules and are presented in Table 5. These theoret- tions involving the Van der Waals type dipole–dipole forces etc.,
ical parameters were calculated in the gas phase, aqueous phase as because the larger the dipole moment the stronger will be the
well as in the protonated form of ALLOX. The calculated parame- intermolecular attraction [86]. The dipole moment of ALLOX is
ters in gas phase as well as in the presence of a solvent did not highest in the protonated form (7.98 Debye (26.614  1030 Cm)),
exhibit important differences. which is higher than that of H2O (l = 6.23  1030 Cm). The high
From Figs. 14 and 15, it could be seen that ALLOX have similar value of dipole moment probably increases the adsorption be-
HOMO and LUMO distributions, which were all located on the en- tween chemical compound and metal surface [88]. Accordingly,
tire alloxazine moiety. This is due to the presence of nitrogen and the adsorption of ALLOX molecules can be regarded as a quasi-sub-
oxygen atoms together with several p-electrons on the entire mol- stitution process between the ALLOX compound and water mole-
ecule. Thus, unoccupied d-orbitals of Fe atom can accept electrons cules at the electrode surface.
from inhibitor molecule to form coordinate bond. Also the inhibitor The local reactivity of the ALLOX was analyzed through an eval-
molecule can accept electrons from Fe atom with its anti-bonding uation of the Fukui indices [89]. These are measurements of the
orbitals to form back-donating bond. Fig. 16 shows the Mulliken chemical reactivity, as well as an indicative of the reactive regions
atomic charges calculated for ALLOX. It has been reported that and the nucleophilic and electrophilic behavior of the molecule.
the more negative the atomic charges of the adsorbed centre, the The regions of a molecule where the Fukui function is large are
more easily the atom donates its electron to the unoccupied orbital chemically softer than the regions where the Fukui function is
of the metal [84]. It is clear from Fig. 16, that nitrogen and oxygen small, and by invoking the HSAB principle in a local sense, one
as well as some carbons atoms carries negative charge centers may establish the behavior of the different sites with respect to
which could offer electrons to the mild steel surface to form a coor- hard or soft reagents. The Fukui function f(r) is defined as the first
dinate bond. Among the nitrogen atoms, the highest negative derivative of the electronic density q(r) with respect to the number
charge is domiciled in N4 while among the oxygen atoms the high- of electrons N at a constant external potential v(r). Thus, using a
est negative charge is found in O2. This shows that the two atoms scheme of finite difference approximations from Mulliken popula-
are the probable reactive sites for the adsorption of iron. tion analysis of atoms in ALLOX and depending on the direction of
Higher values of EHOMO are likely to indicate a tendency of the electron transfer we have [89]:
molecule to donate electrons to appropriate acceptor molecules
with low energy or empty electron orbital. It is evident from Table fkþ ¼ qk ðN þ 1Þ  qk ðNÞ ðfor nucleophilic attackÞ ð24Þ
5 that alloxazine has the highest EHOMO in the gas and aqueous fk ¼ qk ðNÞ  qk ðN  1Þ ðfor electrophilic attackÞ ð25Þ
phase and a lower EHOMO in the protonated form. This means that q ðN þ 1Þ  qk ðN  1Þ
the electron-donating ability of ALLOX is weaker in the protonated fko ¼ k ðfor radical attackÞ ð26Þ
2
form. This confirms the experimental results that interaction be-
tween ALLOX and mild steel is electrostatic in nature (physisorp- where qk is the gross charge of atom k in the molecule i.e. the elec-
tion). The energy of the LUMO is directly related to the electron tron density at a point r in space around the molecule. The N corre-
affinity and characterizes the susceptibility of the molecule to- sponds to the number of electrons in the molecule. N + 1
wards attack by neuclophiles. The lower the values of ELUMO are, corresponds to an anion, with an electron added to the LUMO of
the stronger the electron accepting abilities of molecules. It is clear the neutral molecule; N  1 corresponds to the cation with an elec-
from Table 4 that the protonated form of ALLOX exhibits the lowest tron removed from the HOMO of the neutral molecule. All calcula-
EHOMO, making the protonated form the most likely form for the tions were done at the ground-state geometry. These functions can
interaction of mild steel with alloxazine molecule. Low values of be condensed to the nuclei by using an atomic charge partitioning
the energy gap (DE) will provide good inhibition efficiencies, be- scheme, such as Mulliken population analysis in Eqs. (24)–(26).
cause the excitation energy to remove an electron from the last The calculated Fukui indices for the charged species (N + 1 and
occupied orbital will be low [85]. A molecule with a low energy N  1) as well as the neutral specie (N) are presented in Table 6. For
gap is more polarizable and is generally associated with a high simplicity, only the charges and Fukui functions over the Nitrogen
chemical reactivity, low kinetic stability and is termed soft mole- (N), and oxygen (O) are presented. For a finite system such as an
cule [86]. According to Wang et al. [87], adsorption of inhibitor inhibitor molecule, when the molecule is accepting electrons one
onto a metallic surface occurs at the part of the molecule which has fkþ , the index for nucleophilic attack; when the molecule is
has the greatest softness and lowest hardness. The result from donating electrons, one has fk , the index for electrophilic attack.
It is possible to observe from Table 6 that the nitrogen atoms in
the alloxazine ring (N2, N3 and N4) are the most susceptible sites
Table 5 for electrophilic attacks. These sites present the highest values of
Some molecular properties of alloxazine calculated using DFT at the B3LYP/6-31G
fk which are 0.003 for N1, 0.0028 for N2, and 0.0039 for N4,
(d,p) basis set.
respectively. On the other hand N4, N3 are the most susceptible
Inhibitor EHOMO ELUMO DE l g x sites for nucleophilic attacks. These sites have the highest values
(eV) (eV) (eV) (D)
of fkþ which are 0.01 for N4 and 0.003 for N3, respectively. Further-
ALLOX (gas phase) 6.78 2.74 4.04 4.24 2.02 5.61 more, of the two oxygen atoms in alloxazine, O2 has the highest
ALLOX (aqueous phase) 6.78 2.74 4.04 4.24 2.02 5.61
value of fkþ (Table 6). This result also confirms the work of Kaim
ALLOX (protonated at N2) 11.33 7.45 3.88 7.98 1.94 22.71
et al. [74] that alloxazine bind predominantly through N4 and O2
274 N.O. Obi-Egbedi, I.B. Obot / Corrosion Science 53 (2011) 263–275

Table 6
Calculated Mulliken atomic charges, Fukui functions and philicity indices for heteroatoms of alloxazine using DFT at the B3LYP/6-31G (d,p) basis set.

Atom qN qNþ1 qN1 fkþ fk fko xþk xk


N1 0.568 0.615 0.483 0.047 0.086 0.067 1.067 1.953
N2 0.519 0.642 0.523 0.123 0.003 0.060 2.793 0.068
N3 0.611 0.605 0.614 0.003 0.0028 0.005 0.068 0.006
N4 0.639 0.629 0.600 0.010 0.004 0.015 0.227 0.088
O1 0.455 0.553 0.387 0.098 0.068 0.083 2.225 1.544
O2 0.483 0.565 0.391 0.081 0.092 0.086 1.839 2.089

with a metal ion.Recently, Parr et al. [90], have introduced an elec-


trophilicity index (x) defined as:

v2
x¼ ð27Þ
2g
This was proposed as a measure of the electrophilic power of a mol-
ecule. The higher the value of x, the higher the capacity of the mol-
ecule to accept electrons.
In Eq. (27), v is electronegativity and g is global hardness. For v Fig. 17. Proposed structure of the complex compound formed between alloxazine
and Fe2+.
and g, their operational and approximate definitions are [87]:
IþA
v¼ ð28Þ
2 species (Eq. (34)) which may be adsorbed on the cathodic sites of
IA the mild steel and reduce the evolution of hydrogen:
g¼ ð29Þ
2
ALLOX þ 2Hþ $ ½ALLOXH2þ ð34Þ
According to Koopman’s theorem [91], the energies of the HOMO
and the LUMO orbitals of the inhibitor molecule are related to the The protonated ALLOX, however, could be attached to the mild steel
ionization potential, I, and the electron affinity, A, respectively, by surface by means of electrostatic interaction between SO24 and pro-

the following relation [92]: tonated ALLOX since the steel surface has positive charges in the
acid medium [94]. This could futher be explained based on the
I ¼ EHOMO ð30Þ assumption that in the presence of SO2 4 , the negatively charged
A ¼ ELUMO ð31Þ SO2
4 would attach to positively charged surface. When ALLOX ad-
sorbs on the steel surface, electrostatic interaction takes place by
In order to provide a unified treatment of chemical reactivity partial transference of electrons from the polar atoms (N and O
and selectivity a new concept of philicity has been introduced re- atoms and the delocalized p-electrons around the heterocyclic
cently by Chattaraj et al. [93]. This local philicity index is given as: rings) of ALLOX to the metal surface. In addition to electrostatic
xa ðrÞ ¼ xf a ðrÞ ð32Þ interaction (physisorption) of ALLOX molecules on the steel surface,
molecular adsorption may also play a role in the adsorption process.
or its condensed-to-atom variant for the atomic site k in a molecule A close examination of the chemical structure of alloxazine, re-
is defined as: veals that ALLOX molecules have structure characterized by the
presence of chelation centers mainly located on nitrogens and oxy-
xak ¼ xfka ð33Þ
gen. From theoretical and experimental results obtained, N4 and
where a = +, , o refer to nucleophilic, electrophilic and radical at- O2 are the likely sites of complexation of ALLOX with the Fe2+
tacks, respectively. The xak is capable of providing other local and (Fig. 17) which will result in the formation of a five-membered, re-
global reactivity descriptors. dox active a-iminoketo chelate rings [74]. The UV–Visible absorp-
The electrophilic or nucleophilic power is distributed over all tion spectra of the solution containing the inhibitor after the
atomic sites in a molecule keeping the overall philicity conserved. immersion of the mild steel speciment indicated the formation of
The atomic site with the largest xþk will be the most favourable site
a complex with the steel surface allowing the formation of adhe-
for nucleophilic attack, the highest xk for the electrophilic attack
sive film. Such an adhesive film covered the metal surface isolating
and the highest xok for the radical attack [93]. From Table 5, it is the metal surface from the corrosive media.
clear that x is highest for the protonated form of ALLOX inferring
that in this form, ALLOX has the highest capacity to accept elec- 4. Conclusions
trons from the vacant d-orbital of iron. On the other hand, xþ k is
highest in the N4 and N3 centers (Table 6) signifying that the The following conclusions may be drawn from the study:
nucleophilic attack will take place on these atoms. For all the
atoms of ALLOX, x k is highest on N2 and N3 indicating that these (1) Alloxazine (ALLOX) acts as an inhibitor for the corrosion of
sites would be most favourable for electrophilic attack say mild steel in 0.5 M H2SO4. Inhibition efficiency values
(protonation). increase with the inhibitor concentration but decrease
with rise in temperature suggesting physical adsorption
3.7. Mechanism of inhibition action of alloxazine on mild steel mechanism.
(2) The adsorption of ALLOX on steel surface was found to
From the experimental and theoretical results obtained, we accord with Temkin adsorption isotherm model. The adsorp-
note that a plausible mechanism of corrosion inhibition of mild tion process is spontaneous, exothermic and accompanied
steel in 0.5 M H2SO4 by alloxazine may be deduced on the basis with a decrease in entropy of the system from thermody-
of adsorption. In acidic solutions the inhibitor can exist as cationic namic point of view.
N.O. Obi-Egbedi, I.B. Obot / Corrosion Science 53 (2011) 263–275 275

(3) A first-order kinetics relationship with respect to the mild [35] M. Lebrini, F. Bentiss, H. Vezin, M. Lagrenee, Corros. Sci. 48 (2006) 1291.
[36] I. Ahmad, R. Prasad, M.A. Quraishi, Corros. Sci. 52 (2010) 933–942.
steel was obtained with and without ALLOX from the kinetic
[37] X. Li, S. Deng, H. Fu, G. Mu, Corros. Sci. 51 (2009) 620–634.
treatment of the data. [38] J.O. Olusola, A.K. Oluseyi, O.O. Kehinde, A.O. Olayinka, J.M. Oluwatosin, Port.
(4) UV–Visible spectrophotometric studies clearly reveal the Electrochim. Acta. 27 (5) (2009) 591–598.
formation of Fe-ALLOX complex which may be responsible [39] A. Popova, E. Sokolova, S. Raicheva, M. Christov, Corros. Sci. 45 (2003) 33–58.
[40] S.K. Shukla, M.A. Quraishi, Corros. Sci. 51 (2009) 1007–1011.
for the observed inhibition. [41] A.K. Singh, M.A. Quraishi, Corros. Sci. 52 (2010) 152–160.
(5) Data obtained from quantum chemical calculations using [42] H. Ashassi-Sorkhabi, B. Shaabani, D. Seifzadeh, Appl. Surf. Sci. 239 (2005) 154–
DFT at the B3LYP/6-31G (d,p) level of theory were correlated 164.
[43] A.Y. Musa, A.A.H. Kadhum, A.B. Mohamad, A.R. Daud, M.S. Takriff, S.K.
to the inhibitive effect of alloxazine. Both experimental and Kamarudin, Corros. Sci. 51 (2009) 2393–2399.
theoretical calculations are in excellent agreement. [44] Z. Sibel, P. Dogan, B. Yazici, Corros. Rev. 23 (2005) 217–227.
[45] S.A. Umoren, U.F. Ekanem, Chem. Eng. Commun. 197 (2010) 1339–1356.
[46] A.S. Fouda, A.A. Al-Sarawy, E.E. El-Katori, Desalination 201 (2006) 1–13.
[47] E.A. Noor, A.H. Al-Moubaraki, Mater. Chem. Phys. 110 (2008) 145–154.
References
[48] N.M. Guan, L. Xueming, L. Fei, Mater. Chem. Phys. 86 (2004) 59–65.
[49] X. Li, S. Deng, H. Fu, G. Mu, Corros. Sci. 51 (2009) 1344–1355.
[1] M. Ajmal, A.S. Mideen, M.A. Quraishi, Corros. Sci. 36 (1994) 79–84. [50] M. Sahin, S. Bilgic, H. Yilmaz, Appl. Surf. Sci. 195 (2002) 1–7.
[2] M. Bethencourt, F.J. Botana, J.J. Calvino, M. Marcos, Corros. Sci. 40 (1998) 1803– [51] S.A. Umoren, O. Ogbobe, I.O. Igwe, E.E. Ebenso, Corros. Sci. 50 (2008) 1998–
1819. 2006.
[3] G.O. Santana, L. Sathler, J.A.C.P. Gomes, Evaluation of Corrosion Inhibitors for [52] K.F. Khaled, Appl. Surf. Sci. 230 (2004) 307.
Acidizing Operations, International Corrosion Council, Canada, 2002. [53] M.A. Amin, S.S. Abd El Rehim, M.M. El-Naggar, H.T.M. Abdel-Fatah, J. Mater. Sci.
[4] M. El Achouri, M.R. Infante, F. Izquierdo, S. Kertit, H.M. Gouttoya, B. Nciri, 44 (2009) 6258.
Corros. Sci. 43 (2001) 19–35. [54] M.A. Amin, Q. Mohsen, O.A. Hazzazi, Mater. Chem. Phys. 114 (2009) 908–914.
[5] I.B. Obot, N.O. Obi-Egbedi, N.W. Odozi, Corros. Sci. 52 (2010) 923. [55] L. Tang, X. Li, Y. Si, G. Mu, G. Liu, Mater. Chem. Phys. 95 (2006) 29–38.
[6] I.B. Obot, N.O. Obi-Egbedi, Corros. Sci. 52 (2010) 282–285. [56] Z. Sibel, P. Dogan, B. Yazici, Corros. Rev. 23 (2005) 217.
[7] I.B. Obot, N.O. Obi-Egbedi, Surf. Rev. Lett. 15 (6) (2008) 903–910. [57] G. Moretti, F. Guidi, G. Grion, Corros. Sci. 46 (2004) 387–403.
[8] I.B. Obot, N.O. Obi-Egbedi, Colloids Surf. A Physicochem. Eng. Aspects 330 [58] A.K. Singh, M.A. Quraishi, Corros. Sci. 52 (2010) 1373–1385.
(2008) 207–212. [59] N. Soltani, M. Behpour, S.M. Ghoreishi, H. Naeimi, Corros. Sci. 52 (2010) 1351–
[9] I.B. Obot, N.O. Obi-Egbedi, S.A. Umoren, Int. J. Electrochem. Sci. 4 (2009) 863– 1361.
877. [60] F.M. Donahue, K. Nobe, J. Electrochem. Soc. 112 (1965) 886.
[10] I.B. Obot, N.O. Obi-Egbedi, Corros. Sci. 52 (2010) 198–204. [61] S.V. Ramesh, A.V. Adhikari, Corros. Sci. 50 (2008) 55.
[11] I.B. Obot, N.O. Obi-Egbedi, S.A. Umoren, Corros. Sci. 51 (2009) 1868–1875. [62] E.A. Noor, J. Appl. Electrochem. 39 (2009) 1465–1475.
[12] I.B. Obot, Port. Electrochim. Acta 27 (5) (2009) 539–553. [63] A.M. Badiea, K.N. Mohana, Corros. Sci. 51 (2009) 2231–2241.
[13] I.B. Obot, N.O. Obi-Egbedi, S.A. Umoren, Corros. Sci. 51 (2009) 276–282. [64] I. El Ouali, B. Hammouti, A. Aouniti, Y. Ramli, M. Azougagh, E.M. Essassi, M.
[14] I.B. Obot, N.O. Obi-Egbedi, Corros. Sci. 52 (2010) 657–660. Bouachrine, J. Mater. Environ. Sci. 1 (1) (2010) 1–8.
[15] E.E. Ebenso, H. Alemu, S.A. Umoren, I.B. Obot, Int. J. Electrochem. Sci. 4 (2008) [65] G. Mu, X. Li, G. Liu, Corros. Sci. 47 (2005) 1932–1952.
1325–1339. [66] J.M. Thomas, W.J. Thomas, Introduction to the Principles of Heterogeneous
[16] J. Chastain, D.B. McCormick, Flavin metabolites, in: F. Muller (Ed.), Chemistry Catalysis, fifth ed., Academic Press, London, 1981. pp. 14.
and Biochemistry of Flavoenzymes, vol. 1, CRC Press, Boston, 1991, pp. 196– [67] D.D. Ebbing, S.D. Gammon, General Chemistry, Houghton Mifflin Company,
200. Boston, 2005.
[17] O. Cunningham, M.G. Gore, T.J. Mantle, Biochem. J. 345 (2000) 393–398. [68] O.K. Abiola, J.O.E. Otaigbe, Corros. Sci., 2009, doi:10.1016/j.corsci.2009.07.006.
[18] M. Sikorski, E. Sikorska, I.V. Khemelinskii, R. Gonzalez-Moreno, J.L. [69] O.K. Abiola, Corros. Sci. 48 (2006) 3078–3090.
Bourdelande, A. Siemiarczuk, Photochem. Photobiol. 1 (2002) 715–721. [70] P. Atkins, J. de Paula, A Textbook of Physical Chemistry, seventh ed., University
[19] K. Tatsumi, H. Ichikawa, S. Wada, J. Contam. Hydrol. 9 (1992) 207–212. Press, Oxford, 2002. pp. 873.
[20] I. Ahamad, M.A. Quraishi, Corros. Sci. 52 (2010) 651–656. [71] M.A. Quraishi, M.Z.A. Rafiquee, S. Khan, N. Saxena, J. Appl. Electrochem. 37
[21] A.Y. Musa, A.A. Khadom, A.H. Kadhum, A.B. Mohamad, M.S. Takriff, J. Taiwan, (2007) 1153–1162.
Ins. Chem. Eng. 41 (2010) 126–128. [72] Y. Abboud, A. Abourriche, T. Saffaj, M. Berrada, M. Charrouf, A. Bennamara, N.
[22] A.A. Khadom, A.S. Yaro, A.H. Kadum, J. Taiwan, Ins. Chem. Eng. 41 (2010) 122– Al-Himidi, H. Hannache, Mater. Chem. Phys. 105 (2007) 1–5.
125. [73] Y. Abboud, A. Abourriche, T. Saffaj, M. Berrada, M. Charrouf, A. Bennamara, H.
[23] M. Bouklah, B. Hammouti, M. Lagrenee, F. Bentiss, Corros. Sci. 48 (2006) 2831– Hannache, Desalination 237 (2009) 175–189.
2842. [74] W. Kaim, B. Schwederski, O. Heilmann, F.M. Hornung, Coord. Chem. Rev. 182
[24] S. Chitra, K. Parameswari, C. Sivakami, A. Selvaraj, Chem. Eng. Res. Bull. 14 (1) (1999) 323–333.
(2010) 1–6. [75] H. Chermette, J. Comput. Chem. 20 (1999) 129–154.
[25] A.R. Afidah, J. Kassim, Recent Patents Mater. Sci. 1 (2008) 223–231. [76] R.G. Parr, W. Yang, Density Functional Theory of Atoms and Molecules, Oxford
[26] S.A. Umoren, M.M. Solomon, I.I. Udousoro, A.P. Udoh, Cellulose 17 (2010) 635– University Press, Oxford, 1989.
648. [77] B.S. Jursic, J. Chem. Phys. 104 (1996) 4151–4157.
[27] M.M. Solomon, S.A. Umoren, I.I. Udosoro, A.P. Udoh, Corros. Sci. 52 (4) (2010) [78] R.G. Parr, W. Yang, J. Am. Chem. Soc. 106 (1984) 4049–4050.
1317–1325. [79] R.G. Pearson, J. Am. Chem. Soc. 85 (1963) 3533–3539.
[28] S.A. Umoren, I.B. Obot, E.E. Ebenso, N.O. Obi-Egbedi, Desalination 247 (2009) [80] M. Lashkari, M.R. Arshadi, J. Chem. Phys. 299 (2004) 131–138.
561–572. [81] L.T. Sein, Y. Wei, S.A. Jansen, Comput. Theor. Polym. Sci. 11 (2001) 83–90.
[29] S.A. Umoren, I.B. Obot, N.O. Obi-Egbedi, J. Mater. Sci. 44 (2009) 274–279. [82] E.E. Ebenso, T. Arslan, F. Kandemirli, N. Caner, I. Love, Int. J. Quant. Chem. 110
[30] C. Lee, W. Yang, R.G. Parr, Phys. Rev. B 37 (1988) 785–789. (2010) 1003–1018.
[31] Y. Shao, L.F. Molnar, Y. Jung, J. Kussmann, C. Ochsenfeld, S.T. Brown, A.T.B. [83] G. Gece, S. Bilgic, Corros. Sci., 2010, doi:10.1016/j.corsci.2010.06.015.
Gilbert, L.V. Slipehenko, S.V. Levehenko, D.P. O’Neill, R.A. DiStasio Jr., R.C. [84] K.F. Khaled, S.A. Fadl-Allah, B. Hammouti, Mater. Chem. Phys., 2009,
Lochan, T. Wang, G.J.O. Beran, N.A. Besley, J.M. Herbert, C.Y. Lin, T. Van Voorhis, doi:10.1016/j.matchemphys.2009.05.043.
S.H. Chien, A. Sodt, R.P. Steele, V.A. Rassolov, P.E. Maslen, P.P. Korambath, R.D. [85] G. Gece, Corros. Sci. 50 (2008) 2981–2992.
Adamson, B. Austin, J. Baker, E.F.C. Byrd, H. Dachsel, R.J. Doerksen, A. Dreuw, [86] A. Dwivedi, N. Misra, Der Pharma Chemica 2 (2) (2010) 58–62.
B.D. Dunietz, A.D. Dutoi, T.R. Furlani, S.R. Gwaltney, A. Heyden, S. Hirata, C.P. [87] H. Wang, X. Wang, H. Wang, L. Wang, A. Liu, J. Mol. Model. 13 (2007) 147–153.
Hsu, G. Kedziora, R.Z. Khalliulin, P. Klunzinger, A.M. Lee, M.S. Lee, W.Z. Liang, I. [88] X. Li, S. Deng, H. Fu, T. Li, Electrochim. Acta 54 (2009) 4089–4098.
Lotan, N. Nair, B. Peters, E.I. Proynov, P.A. Pieniazek, Y.M. Rhee, J. Ritchie, E. [89] W. Yang, W.J. Mortier, J. Am. Chem. Soc. 108 (1986) 5708–5711.
Rosta, C.D. Sherrill, A.C. Simmonett, J.E. Subotnik, H.L. Woodcock III, W. Zhang, [90] R.G. Parr, L. Szentpaly, S. Liu, J. Am. Chem. Soc. 121 (1999) 1922–1928.
A.T. Bell, A.K. Chakraborty, D.M. Chipman, F.J. Keil, A. Warshel, W.J. Hehre, H.F. [91] V.S. Sastri, J.R. Perumareddi, Corrosion 53 (1997) 617–622.
Schaefer, J. Kong, A.I. Krylov, P.M.W. Gill, M. Head-Gordon, Phys. Chem. Chem. [92] M.S. Masoud, M.K. Awad, M.A. Shaker, M.M.T. El-Tahawy, Corros. Sci. 52 (2010)
Phys. 8 (2006) 3172. 2387–2396.
[32] M.M. El-Naggar, Corros. Sci. 49 (2007) 2226–2236. [93] P.K. Chattaraj, B. Maiti, U. Sarkar, J. Phys. Chem. A 107 (2003) 4973–4975.
[33] S.A. Umoren, E.E. Ebenso, Mater. Chem. Phys. 106 (2007) 387–393. [94] Y. Li, P. Zhao, Q. Liang, B. Hou, Appl. Surf. Sci. 252 (2005) 1245–1253.
[34] A.Y. El-Etre, Corros. Sci. 45 (2003) 2485–2495.

You might also like